• 検索結果がありません。

The Global Weak Solutions of the Compressible Euler Equation with Spherical Symmetry(Evolution Equations and Nonlinear Problems)

N/A
N/A
Protected

Academic year: 2021

シェア "The Global Weak Solutions of the Compressible Euler Equation with Spherical Symmetry(Evolution Equations and Nonlinear Problems)"

Copied!
28
0
0

読み込み中.... (全文を見る)

全文

(1)

1

The

Global

Weak Solutions of the Compressible Euler

Equation with

Spherical Symmetry

Tetu Makino*, Kiyoshi Mizohata** and Seiji Ukai**

(December, 1991)

Department ofLiberal Arts

Osaka Sangyo University* and

Department of Information Sciences

Tokyo Institute of Technology**

1 Introduction

The compressible Euler equation for an isentropic

gas

in $R^{n}$ is given by

$\rho_{t}+\nabla\cdot(\rho\tilde{u})=0$ ,

(1.1)

$(\rho\vec{u})_{t}+\nabla\cdot(\rho\vec{u}\otimes\vec{u}+p)=0$

,

with the equation ofstate

(1.2) $p=a^{2}\rho^{\gamma}$ ,

where density$\rho$

,

velocity

$\vec{u}$and pressure

$p$ are functionsof $x\in R^{n}$ and $t\geq 0$,

while $a>0$ and $\gamma\geq 1$ are given constants.

Forone dimensional case $(n=1)$

,

the Cauchy problem for (1.1) with (1.2)

has been studied by many authors. Nishida [10] established the existence of global weak solutions, for the first time, for the case $\gamma=1$ with arbitrary

initial data, and Nishida and Smoller [11] for $\gamma\geq 1$ but with small initial

data, both

using Glimm’s

method. DiPerna [3] extended the latter result to

the case of large initial data, using the theory of compensated compactness under the restriction $\gamma=1+2/(2m+1),$ $m\geq 2$ integers. Ding et al

[1], [2] removed this restriction and established the existence ofglobal weak

solutions for $1<\gamma\leq 5/3$

.

数理解析研究所講究録 第 785 巻 1992 年 1-28

(2)

On the other hand, littleis known for the case $n\geq 2$

.

No global solutions

have been known to exist, but only local classical solutions ([5], [6], [8] and

[9]).

In this paper, we will present global weak solutions first

for

the case

$n\geq 2$

.

We will do this, however, only for

the

case of spherically symmetry

with $\gamma=1$

.

As will be seen below, our proof does not work without these

restrictions.

Thus, we look for solutions of the form

(1.3) $\rho=\rho(t, |x|))\tilde{u}=\frac{x}{|x|}\cdot u(t, |x|)$

.

Then, denoting $r=|x|,$ $(1.1)$ becomes

(1.4) $\rho_{t}\rho+_{t}\frac{1}{r^{n-1}+u}(;^{n-}1^{\rho_{P}u_{f}})_{f}==00$ ,

This equation has a singularity at $r=0$

.

To avoid the difficulty caused by

this singularity, we simply deal with the boundary value problem for (1.4)

in the domain $1\leq r<\infty$ (the exterior of a sphere) with the boundary

condition $u(t, 1)=0$, which is identical, under the assumption (1.3), to the usual boundary condition $\tilde{n}\cdot\tilde{u}=0$ for (1.1) where $\tilde{n}$ is the unit normal to the boundary.

Put $\tilde{\rho}=r^{\mathfrak{n}-1}\rho$

.

Then we get from (1.4)

$\tilde{\rho}_{\ell}+(\tilde{\rho}u)_{r}=0$ , (1.5)

$u_{\ell}+u u_{f}+\frac{a^{2}\gamma\tilde{\rho}_{r}}{\tilde{\rho}^{2-\gamma_{\Gamma}\langle n-1)\langle\gamma-1)}}=\frac{a^{2}\gamma(n-1)\tilde{\rho}^{\gamma-1}}{r^{n}\cdot r^{(\mathfrak{n}-1)(\gamma-2)}}$

Introduce the Lagrangean mass coordinates

(1.6) $\tau=t$

,

$\xi=l^{r}\tilde{\rho}(t,r)$ dr.

Then $\xi>0$ as long as $\tilde{\rho}>0$ for $r>1$, and (1.5) is

reformulated

as $\tilde{\rho}_{\tau}+\tilde{\rho}^{2}u_{\xi}=0$ ,

(1.7)

(3)

3

Put $v=1/\tilde{\rho}$ and note that the inverse transformation to (1.6) is

given

by

(1.8) $t=\tau,$ $r=1+ \int_{0}^{\epsilon}v(\zeta,t)d\zeta$

.

Then after changing $\tau$ to $t$ and $\xi$ to

$x,$ $(1.7)$ is written as

$v_{t}-u_{x}=0$ ,

(1.9)

$u_{t}+( \frac{a^{2}}{v^{\gamma}})_{x}$ $\frac{1}{r^{(n-1)(\gamma-1)}}=\frac{a^{2}\gamma(n-1)v^{1-\gamma}}{r^{n}\cdot r^{(n-1)(\gamma-2)}}$

where $r$ is now defined by $r=1+ \int_{0}^{x}v(t, \zeta)d\zeta$

.

Now we restrict ourseves to the case $\gamma=1$

.

Then (1.7) becomes

$v_{t}-u_{x}=0$

,

(1.10)

$u_{t}+( \frac{a^{2}}{v})_{x}=\frac{K}{1+\int_{0}^{x}v(t,\zeta)d\zeta}$

where $K=a^{2}(n-1)$

.

Let us consider the initial boundary value problrm for (1.10) in

$t\geq 0,$ $x\geq 0$ with the following boundary and initial conditions.

(1.11) $u(O, x)=u_{0}(x)$ , $v(O, x)=v_{0}(x)$ ,

for

$x>0$,

(1.12) $u(t, 0)=0$ ,

for

$t>0$.

Let $BV(R_{+})$ denote the space of functions of bounded variation on

$R_{+}=(0, \infty)$

.

Our main result is as follows.

Theorem (Main Result) Suppose that $u_{0}(x),$ $v_{0}(x)\in BV(R_{+})$

,

and that

$v_{0}(x)\geq\delta_{0}>0$

for

all $x>0$ with some positive constant$\delta_{0}$

.

Then (1.10),

(1.11) and (1.12) have a global weak solution which belongs to the class

$u,$ $v\in L^{\infty}(0, T;BV(R_{+}))\cap Lip([0, T];L_{loc}^{1}(R_{+}))$

for

any$T>0$

.

The definition of the weak solution will be given in section 4. This the-orem can be proved by

following

Nishida’s argument [10] based

on Glimm’s

(4)

method. Indeed this can be seen from thefollowing twosimple observations.

First, the homogeneous equation corresponding to (1.10),

$v_{t}-u_{x}=0$ ,

(1.13)

$u_{t}+( \frac{a^{2}}{v})_{x}=0$ ,

is just the same equation as solved by Nishida [10] using Glimm’s method

both on

the

Cauchy problem and the initial boundary value problem. Note

that if $\gamma>1$, the homogeneous equation for (1.9) has a variable coefficient

and hence does not coincide with the one dimensional Euler equation.

The second observation is that, as long as $v\geq 0$, the right hand side of

(1.10),

(1.14) $\frac{K}{1+\int_{0}^{x}v(t,()d\zeta}$

is monotone decreasing in $x$ and has an a priori estimate

(1.15) T.V. $( \frac{K}{1+\int_{0}^{x}v(t,\zeta)d\zeta})\leq K$,

independent of $v$

.

The one dimensional inhomogeneous Euler equation has

been studied in [12]. However, the conditions imposed therein on the

inho-mogeneous

term are not applicable to our (1.14).

These observations allow us to use Nishida’s argument [10] to construct global weak solutions to (1.10), (1.11) and (1.12). More precisely, we will

first construct, in section 2, approximate solutions of the form

{solution

of

Riemann problem

for

(1.13)} $+$

{nonhomogeneous

term}

$\cross t$

.

This is the main idea of [12]. Then in section 3, we will estimate the total variation of the approximate solutions. Thanks to (1.15), this can be done with a slight modification of Nishida’s argument [10]. In section 4, we will show that there exists a subsequence of approximate solutions which

con-vereges

strongly in $L_{loc}^{1}$ for any finite time interval. Finally, for the sake of

completeness, we give in Appendix a detailed proof of two lemmas used in section

3.

These lemmas are due to Nishida [10], but their proofs are not found in the literature.

(5)

5

2 The Difference Scheme

Toconstruct the approximate solutions, we shall usethedifferencescheme developed in [10]. For $l,$$h>0$

,

define

$Y=$ $\{(n , m); n=1,2,3, \cdots, m=1,3,5, \cdots\}$ ,

(2.1)

$A= \prod_{(m,n)\in Y}[\{nh\}\cross((m-1)l, (m+1)l)]$

,

where $l/h$ will be determined later. Choose a point $\{a_{nm}\}\in A$ randomly,

and write $a_{nm}=(nh, c_{nm})$

.

For $n=0$, we put $c_{Om}=ml$

.

We denote

approximate solutions by $u^{l}$ and $v^{l}$

.

Mesh lengths $l$ and $h$ are chosen so that

$l/h>a/( \inf v^{t})$, for anygiven $T>0$

.

We shall show later that there exists a$\delta>0$ such that

inf

$v^{l}\geq\delta>0$

.

For $0\leq t<h,$ $ml\leq x<(m+2)l,$ $m$ : odd, we define

$u^{l}(t,x)=u_{0}^{l}(t, x)+U^{l}(t,x)t$,

(2.2)

$v^{l}(t,x)=v_{0}^{l}(t,x)$,

where $u_{0}^{l}$ and $v_{0}^{l}$ are the solutions of

$v_{t}-u_{x}=0$

,

(2.3)

$u_{\ell}+( \frac{a^{2}}{v})_{x}=0$,

with initial data

$u_{0}^{l}(0, x)=$ (2.4) $\{u_{0}^{0}((m+2)l)u(ml)$, $X>(mX<(m:_{1)l}^{1)l},$’ $v_{0}^{l}(0, x)=\{v_{0}^{0}((m+2)l)v(ml)$

,

$X>(mX<(mI_{1)l’}^{1)l}$ , and (2.5) $U^{l}(t,x)= \frac{K}{1+\Sigma^{\frac{m+1}{j=^{2}1}}v_{0}((2j-1)l)\cdot 2l}$

For $0\leq t<h,$ $0\leq x<l$, we define $u^{l}$ and $v^{l}$ by (2.2)

where $u_{0}^{l}$ and $v_{0}^{l}$

are the solutions of (2.3) with initial boundary data

(6)

(2.7) $u(t, 0)=0,$ $t>0$, and

(2.8) $U^{l}(t,x)=K$

.

Suppose that $u^{l}$ and $v^{l}$ are defined for$0\leq t<nh$

.

For$nh\leq t<(n+1)h$

,

$ml\leq x<(m+2)l,$ $m$ : odd, we define

$u^{l}(t,x)=u_{0}^{l}(t,x)+U^{l}(t,x)$

.

(t–nh),

(2.9)

$v^{l}(t,x)=v_{0}^{l}(t,x)$

,

where $u_{0}^{l}$ and $v_{0}^{l}$ are the solutions of (2.3) with initial data $(t=nh)$

$u_{0}^{l}(nh, x)=$ (2.10) $\{\begin{array}{l}u^{l}(nh-0,c_{nm})u^{l}(nh-0,c_{nm+2})\end{array}$ $x>(m+1)lx<(m+1)l$, $v_{0}^{l}(nh, x)=\{v^{l}(nh-0,c_{nm}^{nm})v_{l}(nh-0,c)_{+2}$ $x>(mx<(m\ddagger^{1)l}1)l$

,

and (2.11) $U^{l}(t,x)= \frac{K}{1+\Sigma^{\frac{m+1}{j=^{2}1}}v^{l}(nh-0,c_{n2j-1})\cdot 2l}$

For $nh\leq t<(n+1)h,$ $0\leq x<l$

,

we define $u^{l}$ and $v^{l}$ as

(2.9) where $u_{0}^{l}$

and $v_{0}^{l}$ are the solutions of (2.3) with initial $(t=nh)$ boundary data

(2.12)$u_{0}^{l}(nh, x)=u^{l}(nh-0, c_{n1}),$ $v_{0}^{l}(nh, x)=v^{l}(nh-0, c_{n1}),$ $x>0$,

(2.13) $u(t, O)=0,$ $t>nh$,

and $U^{l}(t, x)$ is as (2.8).

3 Bounds for Approximate Solutions

System (1.6) is hyperbolic provided $v>0$, with the characteristic roots

and Riemann invariants given by

$\lambda=-\underline{a}$

$r=u+alogv$

,

(3.1)

a’

(7)

7

It is well-known, [10], that all shock wave curves in the (r,s)-plane have the

same figure. (See Figure 1.) Tlie l-shock wave curve $S_{1}$, starting from

$(r_{0}, s_{0})$ can be expressed in the form

(3.2) $s-s_{0}=f(r-r_{0})$ for $r\leq r_{0}$,

and the 2-shock wave curve $S_{2}$ can also be expressed in the form

(3.3) $r-r_{0}=f(s-s_{0})$ for $s\leq s_{0}$,

where

(8)

The l-rarefaction wave curve $R_{1}$ can be expressed in the form

(3.4) $s-s_{0}=0$ for $r\geq r_{0}$,

and the corresponding expression for the 2-rarefaction wave curve $R_{2}$ is

(3.5) $r-r_{0}=0$ for $s\geq s_{0}$

.

Now we must prepare some lemmas to estimate Riemann invariants.

First, let us consider (2.3) with following initial data

(3.6) $u_{0}(x)=\{\begin{array}{l}u_{l}u_{r}\end{array}$ $v_{0}(x)=\{\begin{array}{l}v_{l},x<0v_{f},x>0\end{array}$

Lemma 3.1 Let $u$ and $v$ are the solutions

of

(2.3) and (3.6). Then,

(3.7) $\{r(t,x)\equiv r(u(t,x),v(t,x))\geq ro\equiv\min(r(u,v),r(u,v_{v_{l}^{l}})_{)})_{)}s(t,x)\equiv s(u(t,x),v(t,x))\leq so\equiv\max(s(u^{f_{f}},v^{r_{f}}),s(u_{l}^{l},$

Next consider (2.3) in $t\geq 0,$ $x\geq 0$ with following initial and boundary

conditions

(3.8) $u(O, x)=u_{0}^{+}$ , $v(O, x)=v_{0}^{+}$

,

for

$x>0$

,

(3.9) $u(t, 0)=0$,

for

$t>0$

.

Lemma 3.2 Let $u$ and $v$ are the solutions

of

(2.3), (3.8) and (3.9). Then,

(3.10) $\{\begin{array}{l}r(t,x)\equiv r(u(t,x),s(t,x))\geq r(u_{0}^{+},v_{0}^{+})s(t,x)\equiv s(u(t,x),s(t,x))\leq\max(-r(u_{0}^{+},v_{o}^{+}),s(u_{o}^{+},v_{0}^{+}))\end{array}$

The above two lemmas were proved in [10]. Using thesetwo lemmas, we can

get the following lemma. Lemma 3.3 Let $u^{l}$ and $v^{l}$

be the approximate solutions

defined

in section 2

and put $r_{0}= \min r(u_{0}(x), v_{0}(x))$ and $s_{0}= \max s(u_{0}(x),v_{0}(x))$

.

Then,

for

$0<t<T$,

(9)

9

Let us consider Riemann problem (2.3) and (3.6). Denote by $\Delta r$

(resp $\Delta s$) the absolute value of the variation ofthe Riemann invariant

$r$

(resp s) in the first (resp second) schock wave.

Definition 3.4 We denote

$P(u_{l}, v_{l}, u_{f}, v_{f})=\triangle r+\Delta s$

.

Then we have the following lemma.

Lemma 3.5

(3.12) $P(u_{1},v_{1},u_{3}, v_{3})\leq P(u_{1}, v_{1}, u_{2}, v_{2})+P(u_{2}, v_{2}, u_{3}, v_{3})$,

where $u_{1},$ $u_{2}$ and $u_{3}$ are arbitrary constants and $v_{1},$ $v_{2}$ and $v_{3}$ are arbitrary

positive costants.

We shall prove Lemma 3.5 in the Appendix A.

Denote by $i_{0}^{n\pm}$ the straight line segmentsjoining the points $(0, (n \pm\frac{1}{2})h)$

and $a_{1n}$

.

Let $F(i_{0}^{n\pm})$ be the absolute value of the variation of the Riemann

invariants for all shocks on $i_{0}^{n\pm}$

.

Then we also have the following Lemma.

Lemma 3.6

(3.13) $F(i_{0}^{n+})\leq F(i_{0}^{n-})$

.

This lemma 3.6 will be proved in the Appendix B.

We denote

$Z_{1}=\{l-O, l+O, 3l-0, \cdots , (2m-1)l-0, (2m-1)l+0, \cdots\}$,

$Z_{2}=\{2l, 4l, 6l\cdots 2ml, \cdots\}$

.

Let $Z_{t^{n)}}=Z_{1}\cup Z_{2}\cup\{c_{nm}\}$ and line up the elements $z_{n,i}$ of $Z_{(n)}$ so that

$z_{n,i}\leq z_{n,i+1}$

.

(We regard

$(2m-1)l-0<(2m-1)l+0$

for $m$ : integer. ) Let

$F(nh- O,u^{l}, v^{l})=\frac{1}{2}F(i_{0}^{n-})$

(10)

$F(nh+0, u^{l}, v^{l})= \frac{1}{2}F(i_{0}^{n+})+\sum_{m:odd}P(u^{l}(a_{nm}), v^{l}(a_{nm}),$ $u^{l}(a_{nm+2}),$$v^{l}(a_{mm+2}))$

.

Using Lemma 3.5 and Lemma 3.6, we get

(3.14) $F((n+1)h+0, u^{l}, v^{l})\leq F((n+1)h-0,u^{l}, v^{l})$

.

The followingequality is obvious from the definition of $F,$ $u^{l}$ and $v^{l}$

.

(3.15) $F((n+1)h-0, u_{0}^{l}, v_{0}^{l})=F(nh+0, u^{l}, v^{l})$

.

We also get

$F((n+1)h-0, u^{1}, v^{l})=F((n+1)h-0, u_{0}^{l}, v_{0}^{l})$

$+ \sum_{m:odd}P$(

$u^{l}$($(n+1)h-0$,ml–O),$v^{l}(n+1)h-0$,ml–O),

$u^{l}(n+1)h-0,ml+0),v^{l}((n+1)h-0, ml+0))$.

Lemma 3.7

$P(u^{l}$($(n+1)h-0$

,

ml–O), $v^{l}$( $(n+1)h-0$

,

ml–O),

(3.16) $u^{l}((n+1)h-0,ml+0),v^{l}((n+1)h-0,ml+0)$

$\leq 2h\{U^{l}(nh, (m-1)l)-U^{l}(nh, (m+1)l)\},$ $m$ : odd.

Proof.

From the definition,

$u^{l}$( $(n+1)h-0$,ml-O) $=u_{0}^{l}(nh, ml)+U^{l}(nh, (m-1)l)\cdot h$,

$u^{l}((n+1)h-0, ml+0)=u_{0}^{l}(nh, ml)+U^{l}(nh, (m+1)l)\cdot h$

,

$v^{l}$($(n+1)h-0$,ml–O) $=v^{l}((n+1)h-0, ml+0)=v_{0}^{l}(nh, ml)$

.

Therefore we get

$r^{l}$(

$(n+1)h-0$

,

ml–O) $-r^{l}((n+1)h-0,ml+0)$

(3.17) $=s^{l}$($(n+1)h-0$

,

ml–O) $-s^{l}((n+1)h-0, ml+0)l$

$=h\cross\{U^{l}(nh, (m-1)l)-U^{l}(nh, (m+1)l)\}\geq 0$

Thus thefollowing inequality holds.

(11)

$\int[$

From (3.18), we get (3.16). $\square$

Using Lemma 3.7, we get

$F((n+1)h-0, u^{l},v^{l})-F((n+1)h-0, u_{o}^{l}, v_{0}^{l})$

(3.19)

$\leq 2h\sum_{m:odd}\{U^{l}(nh, (m-1)l)-U^{l}(nh, (m+1)l)\}\leq 2Kh$

From (3.14), (3.15) and (3.19), we get

(3.20) $F((n+1)h+0, u^{l}, v^{l})\leq F(nh+0, u^{l}, v^{l})+2Kh$

Thus we obtain the following lemma. Lemma 3.8

(3.21) $F(nh+0, u^{l}, v^{l})\leq F(+0, u^{l}, v^{l})+2KT\equiv F_{0}+2KT$

Denote by $G(\tau)$ the absolute value of the sum ofnegative variation of$r^{l}$ and

$s^{l}$

for $t=\tau$

.

Then for $nh\leq\tau<(n+1)h$, we get

$($

3.22

$)^{G(\tau)\leq G(nh)+2h\sum_{m:odd}\{U^{l}(nh,(m-1)l)-U^{l}(nh,(m+1)l)\}}$

$\leq G(nh)+2Kh$.

Lemma 3.9

(3.23) $G(nh)\leq 2F(nh+0,u^{l}, v^{l})$

.

Proof.

Denote by $\delta s$ (resp $\delta r$ ) the absolute value ofthe Riemann invariant

$s$ (resp r) in the first (resp second) shock wave. By (3.2) and (3.3),

$\triangle r+\delta s<2\triangle r$ on the first shock and $\delta r+\triangle s<2\triangle s$ on the second shock.

So from (3.17), (3.18) and above arguements, we get (3.23). $\square$

From (3.23), (3.24) and (3.25), for any $\tau(nh\leq\tau<(n+1)h)$,

$G(\tau)\leq G(nh)+2Kh\leq 2F(nh+0, u^{l}, v^{l})+2Kh$

(3.24)

$\leq 2F_{0}+6KT\equiv M_{1}$

.

Now we can establish a priori estimates of $u^{l}$ and $v^{l}$

.

Denote by T.V.

$u$

(12)

Theorem 3.10 For any $T>0$, the variation

of

$u^{l}$ and $v^{l}$ is bounded

uni-formly

for

$h$ and $\{a_{mn}\}$

.

Their upper bound and lower bound, especially the

positive lower bound

of

$v^{l}$, are also uniformly bounded.

Proof.

Denote by $T.V^{+}.u$ (resp $T.V^{-}.u$ ) the absolute value of the positive

(resp negative) variation of$u$

.

Put $f^{l}\equiv 2u^{l}=r^{l}+s^{l}$

.

Then $0\leq f^{l}(t, 0)\leq$

$Kh$

.

Without loss of generality, we assume that $u_{0}(x)$ and $v_{0}(x)$ are constant outside a bounded interval. Let

(3.25) $f^{l}(t, \infty)=r^{l}(t, \infty)+s^{l}(t, \infty)\equiv M_{2}$.

Then from the definition,

$f^{l}(t, 0)+T.V^{+}.f^{\int}$ –T.$V^{-}.f^{l}=f^{l}(t, \infty)$

.

Since T.$V^{-}.f^{l}(t, \cdot)\leq G(t)$ for any $t,$ $(3.24)$ yields

$T.V^{+}.f^{l}=f^{l}(t, \infty)+T.V^{-}.f^{l}-f^{l}(t, 0)\leq M_{1}+M_{2}$

.

Thus we get

(3.26) T.V.$f^{l}=T.V2u^{l}\leq 2M_{1}+M_{2}$

.

From (3.26), we get

$|f^{l}|\leq Kh+2M_{1}+M_{2}\leq KT+2M_{1}+M_{2}\equiv 2M_{3}$

.

Therefore we get

(3.27) $|u_{l}|\leq M_{3}$.

Using Lemma 3.2, we get

$2alogv^{l}=r^{l}-s^{l} \geq r_{0}-(\max(-r_{0}, s_{0})+KT)$

.

Thus we get

(3.28) $v^{l} \geq\exp\frac{r_{0}-(\max(-r_{0},s_{0})+KT)}{2a}\equiv\frac{1}{M_{5}}$ From the definition,

(13)

13

Using Lemma 3.3 and (3.24),

(3.29) $T.V^{+}.r^{l}=-r^{1}(0)+T.V^{-}.r^{l}+r(t, \infty)\leq-r_{0}+M_{1}+r(t, \infty)$

.

In view of (3.27) and (3.29), there exists a positive constant $M_{6}$ such that

(3.30) $v^{l}\leq M_{6}$

$\square$

Theorem 3.11 For any interval $[x_{1}, x_{2}]\subset[O.\infty$), we get

(3.31) $\int_{x_{1}}^{x_{2}}|u^{l}(t_{2}, x)-u^{l}(t_{1}, x)|+|v^{l}(t_{2}, x)-v^{l}(t_{1}, x)|dx$

$\leq M\cdot(|t_{2}-t_{1}|+h)$, $0\leq t_{1},t_{2}<T$,

where $M$depends on $T,$ $x_{1}$ and $x_{2}$, but not on $l$ and $h$

.

Proof.

Without loss ofgenerality, we assume that

$nh\leq t_{1}<(n+1)h<\cdots<(n+k)h\leq t_{2}<(n+k+1)h$

.

Let

$\int_{x_{1}}^{x_{2}}|u^{l}(t_{2},x)-u^{l}(t_{1}, x)|dx$

$\leq I_{1}+I_{2}+\int_{x_{1}}^{x_{2}}|u^{l}(t_{2}, x)-u^{l}((n+k)h+0,x)|+|u^{l}(t_{1}, x)-u^{l}((n+1)h-0,x)|dx$

where

$I_{1}= \int_{x_{1}}^{x_{2}}\sum_{1=1}^{k}|u^{l}((n+i)h+O, x)-u^{l}((n+i)h-O,x)|dx$

$I_{2}= \int_{x_{1}}^{x_{2}}\sum_{1=1}^{k-1}|u^{l}((n+i+1)h-O, x)-u^{l}((n+i)h+O, x)|dx$

and

(14)

Denote by $1_{1^{\alpha,\beta]}}$ the characteristic function of the interval $[\alpha,\beta]$

.

We regard $T.V.-l<x<l=T.V_{0<x<l}$

.

Then,

$I_{1}$

$\leq\sum_{1=0}^{k+1}\sum_{m:integer}\int_{x_{1}}^{x_{2}}T.V_{2ml<x<(2m+2)l}u^{l}((n+i)h-0,x)\cdot 1_{[2ml,(2m+2)l]}dx$ ,

$\leq([\frac{t_{2}-t_{1}}{h}]+2)\cdot(\sup_{0\leq t\leq T}T.Vu^{l}(t, \cdot))\cdot 2l$

.

$I_{2}$

$\leq\sum_{i_{\overline{k}}^{-0}}^{k}\sum_{m}\int_{x_{1}^{x_{2}}}(T.V_{(2m-1)l<x<(2m+1)l}u_{0}^{l}((n+i+1)h-0, x)\cdot 1_{[(2m-1)l,(2m+1)}\eta+Kh)dx$,

$\leq\sum_{:=0}2l$

.

T.$V.u_{0}^{l}((n+i+1)h-0, \cdot)+K(x_{2}-x_{1})h$,

$\leq([\frac{t_{2}-t_{1}}{h}]+1)\cdot(2l\sup_{0\leq t\leq T}T.V.u_{0}^{l}(t, \cdot)+K(x_{2}-x_{1})h)$

.

The remaining terms can be evaluated similarly. For

$\int_{x_{1}}^{x_{2}}|v^{l}(t_{2}, x)-v^{l}(t_{1}, x)|dx$,

we also have a similar estimate. Combining these results gives (3.31). $\square$

4 Convergence of The Approximate Solution

Let $h_{n}=T/n$ and $h_{n}/l_{n}=\tilde{\delta}<\delta\equiv 1/M_{5}$

.

Consider the sequence

$(u^{l_{n}}, v^{l_{n}})(n=1,2, \cdots)$

.

Then from Theorem 3.9 and Theorem 3.10, there exists a subsequence which converges in $L_{loc}^{1}$ to functions (u,v) uniformly for

$t\in[0,T]$

.

Now we shall prove that u(x,t) and v(x,t) are the weak solutions

of initial boundary value problem (1.6), (1.7) and (1.8) provided $\{a_{nm}\}$ is

suitably chosen, namely, they satisfy the integral identity

(4.1)

$\int_{0}^{T}\int_{0}^{\infty}u\phi_{t}+(\frac{a^{2}}{v})\phi_{x}+\frac{K}{1+\int_{0}^{x}v(t,\zeta)d\zeta}\cdot\phi dxdt$

$+ \int_{0}^{\infty}u_{0}(x)\phi(0,x)dx=0$,

(15)

15

for any smooth functions $\phi$ and $\psi$ with compact support in the region

$\{(t, x) : 0\leq t<T, 0\leq x<\infty\}$ and $\phi(t,0)=0$

.

Now we know that $u_{0}^{l}$ and

$v_{0}^{l}$ are weak solutions in each time strip $nh\leq t<(n+1)h$ so that for each

test function $\phi$ satisfying $\phi(t, 0)=0$,

$\int_{nh}^{(n+1)h}\int_{0}^{\infty}u^{l}\phi_{t}+(\frac{a^{2}}{v^{l}})\phi_{x}+U^{l}(t,x)\cdot\phi dxdt$

(43) $+ \int^{\infty}u^{l}(nh+O, x)\phi(nh, x)$

$- \int_{0}^{\infty}u^{\int}((n+1)h-0,x)\phi((n+1)h, x)dx=0$

Ifwe sum this over $n$, weget

$\int_{0}^{T}\int_{0}^{\infty}u^{l}\phi_{t}+(\frac{a^{2}}{v^{l}})\phi_{x}+U^{l}(t,x)\cdot\phi dxdt+\int_{0}^{\infty}u^{l}(0, x)\phi(0,x)$

(4.4)

$=- \sum_{k=1}^{N}\int_{0}^{\infty}\{u^{l}(kh+O, x)-u^{l}(kh-0, x)\}\cdot\phi(kh, x)dx$

where $N=T/h$. When $Narrow\infty$, the right-hand side of the above equality

tends to $0$ for almost every $\{a_{nm}\}\in A$ (see [4]). It is immediate to see that

$\int_{0}^{\infty}u^{l}(0, x)\phi(0,x)dxarrow\int_{0}^{\infty}u_{0}(x)\phi(0, x)dx$ $(Narrow\infty)$

.

Lemma 4.1

(4.5) $U^{l}(t,x) arrow\frac{K}{1+\int_{0}^{x}v(t,\zeta)d\zeta}$ $(Narrow\infty)$

.

locally uniformly

for

$t$ and $x$.

Proof.

Let $nh\leq t<(n+1)h,$ $x\in((m-1)l, (m+1)l),$ $m$ : odd. Then

(4.6) $| \int_{0}^{x}v^{l}(nh, \zeta)d\zeta-\frac{m+1}{\sum_{j=1}^{2}}v^{l}(nh, c_{2j-1n})|\leq\Vert v^{l}||_{\infty}\cdot l$.

On

the other hand

(16)

locally uniformly for $t$ and $x$

.

We get

$| \int_{0}^{x}v^{l}(t, \zeta)d\zeta-\int_{0}^{x}v^{l}(nh, \zeta)d\zeta|$

(4.8) $\leq\int_{0^{\sum_{m:odd}T.V_{(m-1)l<(<(m+1)\iota^{v^{l}(nh,\cdot)\cdot 1_{1(m-1)l,\langle m+1)T}d\zeta}}}}^{x}$

.

$\leq\sup_{0\leq t\leq T}T.Vv^{\int}\cdot 2l$.

From (4.6), (4.7) and (4.8), we get (4.5). $\square$

For each test function $\psi,$ $v^{l}$ also satisfies,

$\int_{0}^{T}\int_{0}^{\infty}(v^{l}\psi_{t}-u^{l}\psi_{x})dxdt+\int_{0}^{\infty}v^{l}(0, x)\psi(O, x)dx$

(49) $=- \sum_{k=1}^{N}\int_{0}^{\infty}\{v^{l}(kh+O, x)-v^{l}(kh-0, x)\}\cdot\psi(kl, x)dx$

$-I_{1}-I_{2}$

.

where

$I_{1}= \sum_{n=0}^{N-1}\int_{nh}^{(n+1)h}U^{l}(t, O)(t-nh)\psi(t, O)dt$

and

$I_{2}= \sum_{n=0}^{N-1}\sum_{m:odd}\int_{nh}^{(n+1)h}\{U^{l}(t,ml+O)-U^{l}(t, ml-0)\}(t-nh)\psi(t,ml)dt$

.

The first term of the the right-hand side of equality (4.9) tends to $0$ for

almost every $\{a_{nm}\}\in A$ (see [4]). It is also immediate to see that

$\int_{0}^{\infty}v^{l}(O, x)\psi(O,x)dxarrow\int_{0}^{\infty}v_{0}(x)\psi(0, x)dx$ $(Narrow\infty)$

.

We shall show that $I_{1},$ $I_{2}arrow 0$ as $Narrow\infty$

.

$I_{1} \leq||\psi\Vert_{\infty}\sum_{n=0}^{N-1}\int_{nh}^{(n+1)h}U^{l}(t, O)(t-nh)dt$

(4.10)

$\leq||\psi\Vert_{\infty}\sum_{n=0}^{N-1}\int_{nh}^{(n+1)h}K(t-nh)dt$

(17)

$t7$

$\sum_{m:odd}\int_{nh}^{\langle n+1)h}\{U^{l}(t,ml+0)-U^{l}(t,ml-0)\}(t-nh)\psi(t,ml)dt\leq K\Vert\psi||_{\infty}h^{2}$.

Thus we get

(4.11) $I_{2} \leq\Vert\psi\Vert_{\infty}\sum_{n=0}^{N-1}Kh^{2}\leq K\Vert\psi\Vert_{\infty}hT$

From above arguments, we can conclude that $u$ and $v$ satisfy (4.1) and

(4.2). Thus we obtain our main result.

Theorem 4.2 (Main Result) Suppose that $u_{0}(x),$ $v_{0}(x)\in BV(R_{+})$ , and

that$v_{0}(x)\geq\delta_{0}>0$

for

all $x>0$ with somepositive constant$\delta_{0}$

.

Then (1.10),

(1.11) and (1.12) have a globalweak solution which belongs to the class

$u,$ $v\in L^{\infty}(O, T;BV(R_{+}))\cap Lip([0, T];L_{loc}^{1}(R_{+}))$

(18)

Appendix

A Proofof Lemma 3.5

Let

$g(x)=-f(-x)$

, and put

$P(u_{1}, v_{1}, u_{2}, v_{2})=\Delta r_{1}+\Delta s_{1}$

$P(u_{2}, v_{2}, u_{3}, v_{3})=\triangle r_{2}+\Delta s_{2}$

$P(u_{1}, v_{1}, u_{3},v_{3})=\Delta r_{3}+\Delta s_{3}$

Then it is obvious that

$\Delta r_{3}+g(\Delta s_{3})+\Delta s_{3}+g(\Delta r_{3})$

$\leq\triangle r_{1}+\Delta r_{2}+\triangle s_{1}+\triangle s_{2}++g(\Delta r_{1})+g(\triangle r_{2})+g(\triangle s_{1})+g(\Delta s_{2})$

We notice that $f^{u}\leq 0$ and hence

$\leq\Delta r_{1}+\Delta r_{2}+\Delta s_{1}+\triangle s_{2}+g(\Delta r_{1}+\triangle r_{2})+g(\triangle s_{1}+\triangle s_{2})$

.

Let $x+g(x)=h(x),$ $\triangle r_{3}=p’,$ $\triangle s_{3}=q’,$ $\triangle r_{1}+\Delta r_{2}=p$ and $\Delta s_{1}+\Delta s_{2}=q$

.

Then

(A.1) $h(p’)+h(q’)\leq h(p)+h(q)$

.

Put

$K=h(p’)+h(q’)$.

We shall estimate $p+q$ from below under the

restriction (A.1). To do this, as $h$ is monotone increasing function, we must

estimate$p+q$ from below under the restriction

(A.2)

$h(p)+h(q)=K$

.

We do this by using Lagrange’s method ofindeterminate coefficients. Put $G(p, q, \lambda)=p+q+\lambda(h(p)+h(q)-K)$

.

Then

$G_{p}=1+\lambda h’(p)=0,$ $G_{q}=1+\lambda h’(q)=0$

.

Because $h^{u}(x)>0$

,

we get $p=q$

.

So

$p+q$ attains its extremum at $p=q$

.

Wecan show that when$p=q,$ $p+q$ is minimumunder the restriction (A2).

Therefore

(19)

19

Hence it follows that

$p=q \geq\frac{p’+q’}{2}$

Thus we get

(A.3) $p+q\geq p’+q’$

.

which proves Lemma

3.5.

B Proof of Lemma 3.6

To prove Lemma 3.6, we must check the following 12 cases:

1) $c_{1n}<l$

,

(1) $S_{2}$ crosses

$i_{0_{n-}}^{n-}$,

(2) $R_{2}$ crosses $\iota_{0}$

$n-$ (3) no wave cross $\iota_{0}$

2) $c_{1n}\geq l$,

(1) $S_{2}$ and $S_{1}$ cross $i_{0}^{n-}$,

(2) $R_{2}$ and $S_{1}$ cross $i_{0}^{n-}$

,

(3) $S_{2}$ and $R_{1}$ cross $i_{0}^{n-}$

,

(4) $R_{2}$ and $R_{1}$ cross $i_{0}^{n-}$,

(5) $S_{1}$ crosses $i_{0_{n-}}^{n-}$, (6) $R_{1}$ crosses $\iota_{n^{0}-}$ , (7) $S_{2}$ crosses $\iota_{0}$ , (8) $R_{2}$ crosses $i_{0}^{n-},.n-$ (9) no wave cross $\iota_{0}$

(20)

Put $r_{+}^{n-1}=r^{l}(a_{1n-1}),$ $s_{+}^{n-1}=s^{l}(a_{1n-1}),$ $r_{-}^{n-1}=-s_{-}^{n-1}$

$=r^{l}((n-1)h+0,0)$, and $\delta_{n-1}=U^{l}(a_{1n-1})$

.

Put $r_{+}^{n-1’}=r^{l}((n-1)h+O, 2l)$ and $s_{+}^{n-1’}=s^{l}((n-1)h+O,2l)$

.

Put $A=(r_{-}^{n-1}, s_{-}^{n-1}),$ $B=(r_{+}^{n-}, s_{+}^{n-1})$ and $B’=(r_{+}^{n-1’}, s_{+}^{n-1’})$

.

Put $C=(r_{+}^{n-1}+Kh, s_{+}^{n-1}+Kh)$

,

(resp $=(r_{+}^{n-1’}+\delta_{n-1}h,$$s_{+}^{n-1’}+\delta_{n-1}h,$$)$ ) if $c_{1n}<l$ (resp $c_{1n}\geq l$).

If$R_{2}$ crosses $i_{0}^{n+},$

$F(i_{0}^{n+_{n+}})=0\leq F(i_{0}^{n-})$, so that it is sufficient to consider the

cases when $S_{2}$ crosses $\iota_{0}$

.

Figure.2 1) $c_{1n}<l$

.

(1) $S_{2}$ crosses $i_{0}^{n-}$ (Figure 2). Denote by I (resp II) the halfspace

$\{(r,s)|r+s<0\}$ (resp $\{(r,s)|r+s\geq 0\}$

.

)

i) $C\in I$

.

In this case $S_{2}$ crosses $i_{0}^{n+}$

.

Denote by V(PQ) the absolute value of the

total variation of$r$ and $s$ by the line segment PQ. From Figure.3, $F(i_{0}^{n+})=V(A’C)\leq V(A’C’)=V(AB)=F(i_{0}^{n-})$

.

(21)

21

Figure.3 ii) $C\in II$

.

In this case $R_{2}$ crosses $i_{0}^{n+}$

.

Then

(B.1) $F(i_{0}^{n-})\geq F(i_{0}^{n+})=0$

.

(2) $R_{2}$ crosses $i_{0}^{n-}$

In this case $B\in II$ so that $R_{2}$ crosses $i_{0}^{n+}$

.

Then

(B.2) $F(i_{0}^{n-})=F(i_{0}^{n+})=0$

.

(3) no wave crosses $i_{0}^{n-}$

In this case $(r_{+}^{n-1}, s_{+}^{n-1})$ is on the line

$r+s=0$

.

Hence $C\in II$

.

It is

(22)

2) $c_{1n}\geq l$

.

(1) $S_{2}$ and $S_{1}$ cross $i_{0}^{n-}$ (Figure.4)

(23)

23

Figure.5 i) $C\in I$

.

From Figure.5,

$F(i_{0}^{n+})=V(A’C)\leq V(A’C’)=V(A^{u}B’)=V(AB’)=F(i_{0}^{n-})$

.

ii) $C\in II$ implies that $R_{2}$

crosses

$i_{0}^{n+}$

.

So

we get (B2).

(24)

(2) $R_{2}$ and $S_{1}$ cross $i_{0}^{n-}$ Figure.6 i) $C\in I$

.

From Figure.6, $F(i_{0}^{n+})=V(A’C)\leq V(A’D)=V(A^{u}E)=V(A^{u}B^{u})$ $\leq V(BB’’)=V(BB’)=F(i_{0}^{n+})$

ii) $C\in II$

.

(25)

25

(3) $S_{2}$ and $R_{1}$ cross $i_{0}^{n-}$

Figure.7

Put $G=$ $(r_{+}^{n-} +\delta_{n-1}h, s_{+}^{n-1}+\delta_{n-1}h)$ and $II=(r^{l}(a_{1n}), s^{l}(a_{1n}))$

.

Then $H$ is on the line $CG$.

i) $H\in I$.

From Figure.7,

$F(i_{0}^{n+})=V(A’H)\leq V(A^{u}G)\leq V(AB)=F(i_{0}^{n-})$

.

ii) $H\in II$, so $R_{2}$ crosses $i_{0}^{n+}$, and we get

$(B2)n-$ (4) $R_{2}$ and $R_{1}$ cross $\iota_{0}$

(26)

(5) $S_{1}$ crosses $i_{0}^{n-}$ Figure.8 i) $C\in I$

.

From Figure.8, $F(i_{0}^{n+})=V(A’C)=V(AE)=V(AD)$ $\leq V(AB’)=F(i_{0}^{n-})$ Thus we get (B1). ii) $C\in II$

.

(27)

27

(6) $R_{1}$ crosses $i_{0}^{n-}$

In this case, it is obvious that $F(i_{0}^{n+})=0$

.

IIence we get (B3).

Cases (7), (8) and (9) are almost the same as cases (1), (2) and (3) in 1).

Thus, we obtain Lemma

3.6.

References

[1] X. Ding,

G.

Chen and P. Luo, Convergence

of

the Lax-Friedrichs scheme

for

isentropic gas dynamics, (I), (II), Acta Math. Sci., 5, (1985),

483-500,

501-540.

[2] X. Ding, G. Chenand P. Luo, Convergence

of

the Lax-Friedrichsscheme

for

isentropic gas dynamics, (III), Acta Math. Sci., 6, (1986),

75-120.

[3] R. DiPerna, Convergence

of

the viscosity method

for

isentropic gas

dy-namics, Commun. Math. Phys., 91, (1983),

1-30.

[4] J. Glimm, Solutions in the large

for

nonlinear hyperbolic systems

of

equations,

Comm.

Pure Appl. Math., 18, (1965),

697-715.

[5] T. Kato, The Cauchy problem

for

quasi-linear symmetric hyperbolic

sys-tems, Arch. Rational Mech. Anal., 58, (1975), 181-205.

[6] P. D. Lax, Hyperbolic systems

of

conservation laws and the mathematical

theory

of

shock waves, SIAM Reg. Conf. Lecture 11, Philadelphia,

1973.

[7] T. P. Liu and J. Smoller, On the vacuum state

for

the isentropic gas

dynamics equations, Advances in Applied Math., 1, (1980),

345-359.

[8] A. Majda, Compressible Fluid Flow and Systems

of

Conservation Laws

in Several Space Variables, Springer-Verlag New York Inc.

1984.

[9] T. Makino, S. Ukai and S. Kawashima,Sur lasolution\‘a support compact

de l’equation d ‘Eulercompressible, Japan J. Appl. Math., 3, (1986),

(28)

[10] T. Nishida, Globalsolutions

for

an initial boundary value problem

of

a

quasilinear hyperbolic system, Proc. Japan Acad., 44, (1968), 642-646.

[11] T. Nishida and J. Smoller, Solutions in the large

for

some nonlinear

hyperbolic conservations, Comm. Pure Appl. Math., 26, (1973),

183-200.

[12] L. A. Ying and

C.

H. Wang,

Global

solutions

of

the Cauchy problem

for

a nonhomogeneous quasilinear hyperbolic system, Comm. Pure Appl.

参照

関連したドキュメント

It is suggested by our method that most of the quadratic algebras for all St¨ ackel equivalence classes of 3D second order quantum superintegrable systems on conformally flat

Theorem 4.2 states the global existence in time of weak solutions to the Landau-Lifshitz system with the nonlinear Neumann Boundary conditions arising from the super-exchange and

Sun, Optimal existence criteria for symmetric positive solutions to a singular three-point boundary value problem, Nonlinear Anal.. Webb, Positive solutions of some higher

In this work, we present a new model of thermo-electro-viscoelasticity, we prove the existence and uniqueness of the solution of contact problem with Tresca’s friction law by

For the three dimensional incompressible Navier-Stokes equations in the L p setting, the classical theories give existence of weak solutions for data in L 2 and mild solutions for

In Section 7, we state and prove various local and global estimates for the second basic problem.. In Section 8, we prove the trace estimate for the second

Subsequently, Xu [28] proved the blow up of solutions for the initial boundary value problem of (1.9) with critical initial energy and gave the sharp condition for global existence

[CFQ] Chipot M., Fila M., Quittner P., Stationary solutions, blow up and convergence to sta- tionary solutions for semilinear parabolic equations with nonlinear boundary